1. Introduction
This study focuses on nonlinear equations that incorporate the nonlocal parabolic Monge–Ampère equations, which are defined as
and therein, we assume that the function
f belongs to the space
, with
for any
satisfying
; the term
denotes the Cauchy principal value. Here,
A represents a positive definite square matrix, denoted as
. The constraints
and
collectively imply that
is bounded above by
. Further details regarding the nonlocal Monge–Ampère operator
can be found in [
1].
For the integral presented in (1) to be well defined, we impose the condition that
u belongs to the space
and is locally
, where
The space
is the space of functions
such that the semi-norm is
for some and thus for all
. This space is well suited for nonlocal fractional operators because it controls the growth and the regularity of the function in a way that is compatible with the nonlocal nature of the integral in the definition of
.
The space
consists of functions that are locally
and whose first-order derivatives are locally Lipschitz continuous. That is, for any compact set
, there exists a constant
such that for all
,
For a function , the integral (2) is well defined. The local regularity ensures that the integrand behaves nicely in a neighborhood of x. The condition guarantees that the integral converges at infinity.
The form of the nonlocal Monge–Ampère operator presented in Equation (1) is motivated by several deep connections in analysis, probability, and differential geometry. From theoretical motivation, the classical Monge–Ampère equation arises in optimal transport and differential geometry, where it relates the determinant of the Hessian of a convex function to a given measure. The nonlocal version in Equation (1) extends this by replacing the local Hessian with a nonlocal fractional operator, inspired by the fractional Laplacian
, with the following relationship:
represents the fractional Laplacian, which is defined here as
where
; the normalization constant
in (3) arises from the Fourier transform definition of the fractional Laplacian and ensures consistency with the standard Laplacian as
. The explicit expression for
is
stands for the Cauchy principal value, which is defined as
where
is the ball of radius
centered at
x. This limit excludes a small neighborhood around the singularity and then takes the limit as the neighborhood shrinks to zero. More explanations for the normalization constant
and Cauchy principal value
about the fractional Laplacian can be found in [
2,
3]. Equations featuring the fractional Laplacian (3) have been extensively studied by numerous researchers, as evidenced in [
4,
5,
6,
7,
8,
9,
10] and the cited literature. The nonlocal nature of the fractional Laplacian introduces substantial challenges in its analysis. To address these challenges, the method of moving planes has been utilized to explore the qualitative characteristics of solutions to equations involving nonlocal operators, as referenced in [
11,
12,
13].
In this study, a direct sliding method is utilized to handle the nonlocal Monge–Ampère operator. The well-known sliding method was initiated by Berestycki and Nirenberg [
14,
15,
16] to analyze the qualitative properties of positive solutions for local elliptic equations. Subsequently, Wu and Chen [
17] refined this approach into a direct sliding method, which has demonstrated its utility in various applications, including deducing monotonicity, one-dimensional symmetry, uniqueness, and the nonexistence of solutions for elliptic equations and systems involving fractional Laplacians and
p-Laplacians. For further details, refer to [
18,
19,
20,
21], with an extensive survey available in [
22]. This direct method circumvents the necessity for the classical extension techniques outlined in [
4] and addresses challenges posed by the nonlocality of fractional operators. For example, the direct method of moving planes relies on local comparisons (e.g., maximum principles for local PDEs); using the sliding method, one can slide
u and use the fact that
u decays at infinity to derive monotonicity or symmetry. The direct method of moving planes requires constructing auxiliary functions (e.g.,
) and analyzing their critical points, while the sliding method often reduces to showing that
u is monotone in a certain direction. For
u defined on
, sliding
u in the
direction and comparing
with
is more straightforward. The direct method of moving planes typically requires the domain to be bounded or symmetric to apply the maximum principle; for unbounded domains (e.g.,
), the method may not directly apply, and the direct sliding method naturally extends to unbounded domains by using the decay properties of
u at infinity. Furthermore, the direct sliding method can be utilized to extend and validate Gibbons’ conjecture within the context of other fractional elliptic equations that incorporate various nonlocal operators (see [
23,
24,
25,
26,
27]).
In [
28], Du and Wang employed the method of moving planes to establish the monotonicity of positive solutions for nonlocal parabolic Monge–Ampère equations as
in the half-space. In this study, the direct sliding method is applied to demonstrate the monotonicity of positive solutions for nonlocal parabolic Monge–Ampère equations, which are defined as
in the whole space. When calculating (1), the integrand
is well behaved at infinity due to the condition
. The local
regularity of functions in
ensures that the function
u and its first-order derivatives are well behaved in any compact subset of
. This is important for the time derivative term
. In [
28], Du and Wang demonstrated that the fractional Monge–Ampère operator is strictly elliptic, which allowed them to apply the well-known regularity results for uniformly elliptic operators. In contrast to the typical research focusing on the regularity aspects of fractional Monge–Ampère operators such as in [
29], this work delves into the maximal regularity properties of such operators. To accomplish this objective, the methodology introduced by Liu (2024) in [
30] is employed, which entails ensuring that the norm of the spatial operator remains below 1. Nevertheless, owing to the nonlocal characteristic of the Monge–Ampère operator, the primary hurdle resides in devising strategies to enforce this norm constraint. Traditionally, researchers might posit the uniform convergence and boundedness of the solution
u to address such challenges. In this article, it is precisely this approach that has been employed.
Section 3 provides a more detailed exposition.
2. Basic Setup and Main Results
The present work adopts a strategy involving the estimation of singular integrals associated with the nonlocal parabolic Monge–Ampère equations, which are conducted along a sequence of points approximating the maximum. The inspiration for these ideas primarily stems from the research conducted in [
31], our objective is to establish the subsequent theorem.
Theorem 1. Let be a solution ofwith conditionand with assumptionSuppose that the function is continuous and exhibits nonincreasing behavior in the vicinity of . Additionally, assume that u is uniformly continuous. Under these conditions, u must be strictly monotonically increasing with respect to the variable , and it depends solely on . Let us denote the vector
x as
For every
belonging to the set of real numbers
, we define
Consider
as a bounded region within
, exhibiting convexity along the
axis. By sliding
downward
units, we obtain
, which is shown in
Figure 1:
Define
and
Suppose
constitutes a positive solution to the equation given by (
4). We then proceed to compare the magnitudes of
with
Let
Our proof is structured into two distinct phases:
The first step: Begin sliding downward by τ units along the axis.
The work shall demonstrate that, provided that
is sufficiently near to
, or equivalently when
attains a sufficiently large magnitude, the set
exhibits a narrow characteristic, which implies that
The second step: Decrease τ as long as holds to its limiting position.
This work aims to demonstrate that the limiting position corresponds to . In the second step, the proof will be divided into two cases: one is , and the other is . In both cases, it will be shown that the limiting position is . After the second step has been completed, it will be proven that , . Consequently, this concludes the demonstration of the monotonicity property for the solutions of parabolic Monge–Ampère equations across the entire space. The concluding section will establish that the function is solely dependent on the variable .
3. Maximal Regularity of Nonlocal Parabolic Monge–Ampère Equations
This section utilizes the theorem proposed by Liu (2025) [
10] to substantiate the existence of solutions for nonlocal parabolic Monge–Ampère equations.
Theorem 2. Let , and assume that is a classical solution ofand then, the solution of (7)
satisfies the - maximal regularity estimate:for any . This is because , and is dense in . In distribution theory, even if a function is not differentiable in the classical sense, its derivatives can still be defined as distributions or generalized functions. For a
function
u, its first-order derivatives
are Lipschitz continuous, which implies they are absolutely continuous and differentiable almost everywhere. Although
may not be differentiable at isolated points, we can define
as a distribution derivative or a weak derivative. If for all test functions
, the following equation
holds, then for a
function
u, the existence of such
can be proven, and it is related to the Lipschitz continuity of
. Since
is Lipschitz continuous, there exists a constant
L such that for all
, we have
This continuity ensures that the variation of
is bounded, allowing us to define its distribution derivatives.
Maximal regularity is a powerful tool in the theory of partial differential equations. It provides a quantitative way to measure the regularity of solutions of PDEs in terms of the regularity of the right-hand side of the equation. While the basic well-posedness results (existence, uniqueness, and continuous dependence on f) only tell us that a solution exists and is unique for a given f, the maximal regularity estimate provides a precise quantitative relationship between the solution and the right-hand side. It tells us how the regularity of the solution is controlled by the regularity of f. For example, if f is in a certain - space, then we know exactly how the - norms of and are bounded in terms of the - norm of f.
The demonstration of Theorem 2 is grounded in the proof of the maximal regularity for parabolic fractional Laplacian equations as presented in [
10]. The immediate priority is to approximate the form of the Monge–Ampère operator (1) to the form of the fractional Laplacian (3) in a certain way.
In an isotropic case where
(the identity matrix), the nonlocal Monge–Ampère operator simplifies to
This is precisely the fractional Laplacian
, evaluated at a fixed time
t.
For general matrices
A with
, the nonlocal Monge–Ampère operator incorporates anisotropic effects. The transition between nonlocal Monge–Ampère operators and fractional Laplacians can be explored through the lens of the Riesz fractional derivative. The Riesz fractional derivative is a well-known alternative definition of the fractional Laplacian. For a function
, the Riesz fractional derivative of order
in
n-dimension is defined in terms of the Fourier transform. If
is the Fourier transform of
, the Riesz fractional derivative
is given by
where
, and
. In Fourier space, the Riesz fractional derivative corresponds to multiplication by
. For the nonlocal Monge–Ampère operator, the Fourier symbol would involve an infimum over all possible "anisotropic" symbols of the form
, where
A is a matrix with
:
The Fourier transform of homogeneous equation
is
and we derive
for some constant
and take
as the eigenvalue so that (11) forms a semi-group. Then, the process of finding the maximal regularity of the nonlocal parabolic Monge–Ampère problem could refer to the nonlocal parabolic fractional Laplacian equations in [
10]. The proof hinges on leveraging the
-
maximal regularity properties of solutions to heat equations. This approach is justified because the Fourier transforms associated with heat equations exhibit structural parallels to those of fractional Laplacians. Furthermore, fractional Laplacians share a functional form akin to the Riesz fractional derivative, which bears resemblance to the nonlocal Monge–Ampère operators. A comprehensive account of the proof is provided in Theorem 1 of [
10].
In the reference [
30], Liu succinctly outlined the fundamental logic and principles dictating the existence of maximal regularity for both parabolic and hyperbolic differential equations. Based on the information in this reference, we can substantiate that the prerequisite for achieving maximal regularity in parabolic differential equations is that the eigenvalues of the operator corresponding to spatial variables are less than 1. In (8), we observe that as
and
,
could be bounded by 1; yet, when
is not adequately large or
t takes a negative value, the following criterion needs to be satisfied to ensure that the eigenvalues of the nonlocal Monge–Ampère operator remain less than 1:
The condition (12) implies that
, and combining this with
, we have
which leads to
and is equivalent to stating that
To satisfy (12) and (13), we might use
norm (maximum value) bounds on
to control the eigenvalues of
; for example, in (
5) and (
6), we have
and
to control the growth of
at different points, thereby regulating the eigenvalues of the nonlocal Monge–Ampère operator (9). With the above conditions on
u, by employing the sliding technique, we deduce the monotonic behavior of the solutions pertaining to nonlocal parabolic Monge–Ampère equations throughout the entire space
.
4. Monotonic Characteristics of Solutions Within
This section elaborates on the comprehensive demonstration of Theorem 1 utilizing the sliding method. Our proof is structured into two distinct stages.
Step 1. It is the aim to demonstrate that, provided
attains a sufficiently large value,
Otherwise, if (16) is not satisfied,
and consequently, there is a sequence
satisfying the condition that
Let
. Given that
approaches
A as
tends to
, it follows that there exists a positive constant
satisfying
Given
such that
So,
. Set
According to (
17), there exists a sequence
converging to 0, where
Set
Given that we possess
and
Subsequently, there is a pair
fulfilling the condition that
Therefore,
This is because we have
Therefore,
According to the definition provided for
, it follows that
We also have
The final inequality at the lower end is valid, owing to .
Therefore,
in which
c represents a positive constant associated with
.
Upon considering scenarios where attains a sufficiently large value, the following two possibilities emerge:
The value of approximates 1;
The value of approximates .
Given the inequality
in the first scenario, both
and
are close to 1. Conversely, in the second scenario, both
and
are close to
. Consequently, regardless of the scenario, we can leverage the monotonicity property of
f to establish that
Consequently, it follows that
Thus, we derive
A contradiction is reached, thereby confirming the validity of (
16).
Step 2. The inequality (
16) serves as an initial condition, enabling us to initiate the sliding process. Subsequently, we progressively reduce the value of
until
is satisfied at its limiting position. We then define
Within this section, our objective is to demonstrate that
.
Otherwise, we have
. Initially, this work establishes the proof for
Should (
19) fail to be satisfied, it follows that
Then, a sequence
exists, where each element is contained within
, satisfying the following condition:
Denote
. Let
be defined as
So,
. Set up
Consequently, a sequence
can be identified, which converges to 0, fulfilling the condition that
Set
Given that
it can subsequently be established that there exists
in
, satisfying
Analogous to the demonstration provided in Step 1, from one perspective, we observe that
Conversely, in accordance with the definition of
, it follows that
Let
Given that the function
u exhibits uniform continuity, we can invoke the Arzelà–Ascoli Theorem to conclude that
As
k approaches infinity, leveraging the continuity of the function
f and drawing upon the results from Equations (
21) and (
22), we arrive at the conclusion that
It follows that
Given that the sequence
is bounded, we obtain that
For every natural number
m, it holds that
Choose
to be sufficiently negative and
m to be sufficiently large. Under these conditions, we observe that
approaches
, while
approaches 1. This scenario leads to a contradiction. Consequently, Equation (
19) must hold true.
Subsequently, this work shall demonstrate the existence of a positive constant
satisfying
Based on Equation (19), we can assert the existence of a sufficiently small
for which
At this point, it suffices to establish the proof for the case where
as
If (25) does not hold, then
We can identify a sequence
satisfying
, where
. Given the conditions, it follows that the sequence
is bounded. Define
. Then, there exists a sequence
with
, along with points
and
, such that
Similarly to previous steps, on the one hand, it holds that
On the other hand,
where
.
As
k tends to infinity, we observe that
approaches 0. This leads to a contradiction, and we consequently arrive at Equation (
23), which is in conflict with the definition of
. Hence, we deduce that
.
To conclude, the work shall demonstrate that the function u is strictly increasing with respect to the variable and that depends solely on .
It has been established that
Suppose that there exists a point
in
satisfying
. In such a case,
constitutes the maximum point of the function
over the domain
. Consequently,
On the one hand,
On the other hand,
The penultimate inequality is valid because is nonpositive across , and specifically negative within the neighborhood .
As
approaches 0, it follows that
This leads to a contradiction. Consequently, we obtain
Subsequently, this work aims to demonstrate that the function is solely dependent on the variable .
Should we substitute
with
, the reasoning remains valid following the preceding methodology. Here,
represents an arbitrary upward-pointing vector with
. By employing analogous reasoning to that utilized in Step 1 and Step 2, we can deduce that, for every such vector
,
As
approaches 0, and leveraging the continuity of the function
u, we can conclude that for any arbitrary vector
satisfying
,
By replacing
by
, we can obtain that for arbitrary
with
,
Equation (
26) implies that the function
u does not depend on the variables
. As a consequence, we can express
as
.
This establishes the validity of Theorem 1.